AM 258

- Computational Fluid Dynamics -

Instructor: Professor George Karniadakis



CONTENTS

(Chapter :1 - Chapter: 9)

Chapter 1:
Introduction (2 lectures)
$\bullet$ Navier-Stokes: (u,p), $(u,\omega)$, Conservative Form, Approximations, B.C., Mathematical Nature, Computational Complexity.
Chapter 2:
The Finite Difference Method (4 lectures)
$\bullet$ Taylor Expansions, Difference Operators, Explicit Formulas, Fonberg's Construction, Implicit Formulas, Lele's Approach, Example, Thomas's Algorithm for Periodic Matrices, Two-dimensional Stencils.
Chapter 3:
Time-Stepping Algorithms (2 lectures)
$\bullet$ IVP, Consistency and Stability, Multi-step Methods (Derivation, Stability Regions), Polynomial Stability Methods.
Chapter 4:
Parabolic Equation (2 lectures)
$\bullet$ Explicit/Implicit Schemes, Discrete Perturbation Analysis, von Neumann Analysis, Matrix Methods, Spectra of Diffusion Operator, Iterative Methods of Elliptic Equations (Jacobi, Gauss Seidel, $\ldots$).
Chapter 5:
Hyperbolic Equations
$\bullet$ Linear Advection Equations, Review of Basic Properties of PDEs, Euler-Forward/Center-Difference Scheme, Euler-Forward Upwind-Differencing Scheme, Lax-Friedrichs Scheme, Second-Order Upwind Scheme, Lax-Wendroff Scheme, Crank-Nicolson/Center Differencing Scheme, Adam-Bashford/Center-Differencing Scheme, Effects of Boundary Conditions, Finite Element Discretizations.
Chapter 6:
Convection-Diffusion Equation
$\bullet$ The Peclet Number (Physical), Crank-Nicolson/Center Difference Scheme, Euler-Forward/Center Difference: Stability Condition, The (Numerical) Peclet Number Controversy; Mixed Explicit Schemes, Semi-Implicit Schemes, Boundary Conditions Normal Boundary Layers.
Chapter 7:
Vorticity-Based Formulations
Vorticity Streamfunction Formulation $(\Psi - \omega)$: Governing Equations, Boundary Condition (Thom's Pearson's, Woods's) 1D Model $\Psi - \omega$ Formulation: Explicit Euler, Semi-Implicit, Full-Implicit; Green's Functions; Vorticity, Velocity Formulations: Green's Functions for 1D Model Problem, Penalty Method; Application in 2D: Flow Over a Backward Facing Step, Streamfunction-Only Formulation.
Chapter 8:
Navier-Stokes in Primitive Variables
Governing Equations-Pressure Interpretation, Stokes Model Problem, Staggered Mesh/Implicit Discretization: Discrete Poisson Equation, Green's Functions Implementation, MAC Grid; Non-Staggered Mesh (Implicit Discretization): Consistent 2D Poisson Equation, Discretization of Stoke Model Problem, Pressure Boundary Conditions, Classical Splitting Scheme, High-Order Splitting Scheme, Other Splitting Schemes; Staggered Mesh (Explicit Discretization), The Artificial Compressibility Methods: Basic Idea, MAC Discretization, The Choice of Parameter C, The Turkel Preconditioner, Stability of Artificial Compressibility Method, The Original Chorin Scheme.
Chapter 9:
Examples/Applications
Chapter 1: Introduction
Incompressible Navier-Stokes Equations


I. Velocity-Pressure Formulation (primitive variables):
\begin{figure}
\centerline{\psfig{file=lec1.2.eps,width=3in}}
\end{figure}

Eulerian description

\begin{displaymath}\hspace*{1in} \left \{ \begin{array}{llll}
\displaystyle{\fr...
...{\rm constrained set of PDEs/mixed IVP/BVP}}\end{array}\right .\end{displaymath}

where

\begin{displaymath}\tau_{ij} = \mu (\frac{\partial u_j}{\partial x_i} + \frac{\partial u_i}{\partial x_j})\end{displaymath}

is the stress tensor for incompressible fluids.
*
Pressure is not a thermodynamic quantity, i.e. $p=C\rho^{\gamma}$ is not valid. It is a constraint, projecting the solution $\vec{u}(x,y,z,t)$ onto a divergence-free space. Otherwise it would overdetermine the system. Note that $\sqrt{\frac{dp}{d\rho}} \rightarrow \infty$ (sound speed). Therefore, disturbances propagate infinitely fast in incompressible media.
*
Viscous terms (assuming linear Newtonian shear stress law): It may be that $ \nu =\mu/\rho = \nu (x,y,z,t)$, then the full stress tensor should be used instead of $\nu \nabla^2 \vec{u}$, i.e. $\nabla \cdot [\nu(\nabla \vec{u} + \nabla^{T} \vec{u})]$. For example, in LES formulations, $(\nu = \nu_{0} + \nu_{e})$, or temperature-dependent viscosity in stability studies.
*
Convective/Advective Terms: $\frac{D\vec{u}}{Dt}$: = Note that the continuum forms are all equivalent but semi-discrete forms may differ.
* Semi-discrete forms for Non-Linear Terms in incompressible Navier-Stokes:

Ref: ``On the rotation and skew-symmetric forms for incompressible flow simulation,'' T.A. Zang, Appl. Num. Math.

For $\nu = 0$ (inviscid, frictionless) flows and no source term $(\vec{F} = 0)$, the N-S conserve linear momentum $\int u\; d\; \Omega$ and kinetic energy $\int\frac{1}{2} \vert u\vert^{2}d \Omega $. However, for semi-discrete system (i.e. continuous in time but discrete in space) we have: *Computational Complexity:
-
Rotation: 6 derivatives
-
Skew- symm: 18 derivatives
-
Convective/divergence/alternate: 9 derivatives
*Conservation Property of Skew-Symmetric Form:

\begin{displaymath}\frac{\partial \phi}{\partial t} + {\cal L}_{N} \phi = 0
\Rightarrow (\phi,\phi_{t})+ (\phi,{\cal L}_{N} \phi) = 0\end{displaymath}

But
\begin{figure}
\centerline{\psfig{file=lec1.3.eps,width=3in}}
\end{figure}


\begin{eqnarray*}(\phi, {\cal L}_{N}\phi) = - ({\cal L}_{N}\phi, \phi) = 0 & \Ri...
...}) = 0\\
& \Rightarrow & \int \phi^{2} d\Omega = {\rm const.}
\end{eqnarray*}




Note: Skew symmetric matirx

\begin{displaymath}A = - A^T \;(n\times n)\end{displaymath}



Note: If n = odd $\Rightarrow$

\begin{displaymath}\Rightarrow \vert A\vert = 0\end{displaymath}



Review: The general Navier-Stokes equations are:

\begin{displaymath}\rho \frac{D\vec{u}}{Dt} = -\nabla p + \nabla \cdot \stackrel{=}{\tau} + \rho \vec{F}\end{displaymath}

Newtonian:

\begin{displaymath}\stackrel{=}{\tau}_{ij} = \mu \left (\frac{\partial u_{j}}{\p...
...x_{j}}\right ) +
\lambda (\nabla \cdot \vec{u}) \delta _{ij}\end{displaymath}

Stokes hypothesis: $2\mu + 3\lambda = 0$, from experimental evidence (equilibrium). Thus,

\begin{displaymath}\stackrel{=}{\tau}_{ij} = \mu \left[ \left (\frac{\partial u_...
...ight ) - \frac{2}{3} (\nabla \cdot \vec{u}) \delta_{ij}\right ]\end{displaymath}

constant $\mu$:

\begin{displaymath}\nabla \cdot \stackrel{=}{\tau} = \mu \left [\nabla^{2} \vec{...
...{u})\right ] = \mu
\nabla^2 \vec{u}, \; {\rm (incompressible)}\end{displaymath}



* Viscoelastic fluids: $\stackrel{=}{\tau} = \tau^{p} {\rm (polymeric) + \tau^{s}}\;{\rm (Newtonian)}$; $\tau^{s} = \mu (\nabla u + \nabla^{T} u)$

In Non-Newtonian fluids (e.g. blood flow), we then need constitutive equations for $\tau^{p}$, the polymeric contributions, which are typically of hyperbolic nature, e.g.,

\begin{displaymath}\lambda \left \{ \frac{\partial \tau^p}{\partial t} + \vec{u}...
...}\right \} = \mu_{p}
[ \nabla \vec{u} + \nabla^{T} \vec{u}],\end{displaymath}

where $\mu_p$ is the polymeric shear viscosity.

II. Velocity-Vorticity Formulation: $\vec{\omega} = \nabla \times \vec{u}$

\begin{displaymath}\frac{D\vec{\omega}}{Dt} = (\vec{\omega}\cdot \nabla)\vec{u} + \nu \nabla^{2} \vec{\omega}\end{displaymath}



or

\begin{displaymath}(*) \left \{ \begin{array}{lll}
&\displaystyle{\frac{\partia...
...l \Omega} \vec{u} \cdot \hat{n} \; ds = 0}
\end{array}\right .\end{displaymath}

System (*) is equivalent to the $(\vec{u},p)$ formulation. The proof, for simply-connected domains, is done in 3 steps.
1.
$\nabla \cdot \vec{\omega} = 0\; {\rm in}\; \Omega$?
2.
$\vec{\omega} = \nabla \times \vec{u} \; {\rm on}\; \partial \Omega \Rightarrow \vec{\omega}
= \nabla \times \vec{u} \; {\rm in}\; \Omega$
3.
$\nabla\cdot \vec{u} = 0\; {\rm in}\; \Omega$?

\begin{displaymath}\mbox{{\rm Let us define}}\; Q \equiv \nabla \cdot \omega \ri...
...rrow \frac{\partial Q}{\partial t} = \nu \nabla^{2} Q\leqno{1.}\end{displaymath}

on

\begin{displaymath}\partial \Omega \colon \; Q = \nabla \cdot \omega =\nabla \cd...
...\times \vec{u}) = 0 \Rightarrow Q \equiv 0 \; {\rm in}\; \Omega\end{displaymath}


\begin{displaymath}\nabla^{2} \vec{u} = - \nabla \times \vec{\omega} = \nabla (\...
...cdot \vec{u}) - \nabla \times (\nabla \times \vec{u})\leqno{2.}\end{displaymath}


\begin{displaymath}\rightarrow \nabla \times [\vec{\omega} - \nabla \times \vec{...
...imes \vec{u}] = - \nabla \times \nabla (\nabla \cdot \vec{u})=0\end{displaymath}


\begin{displaymath}\rightarrow \nabla \{\nabla \cdot [\vec{\omega} - \nabla \tim...
...vec{u}]\} - \nabla^{2} [\vec{\omega} - \nabla \times \vec{u}]=0\end{displaymath}


\begin{displaymath}\left . \begin{array}{ll}
\Rightarrow \nabla^{2} (\vec{\omeg...
...arrow \vec{\omega} = \nabla \times \vec{u}\; {\rm in}\; \Omega \end{displaymath}


\begin{displaymath}\left .\begin{array}{lll}
\nabla^{2} \vec{u} & = &\nabla (\n...
...tarrow \nabla \cdot \vec{u} = 0\; {\rm in} \; \Omega.\leqno{3.}\end{displaymath}



III. Non-Dimensionalization:
\begin{figure}
\centerline{\psfig{file=lec1.4.eps,width=3in}}
\end{figure}

We define $ \bar{u} = \frac{u}{u_{0}}; \; \bar{x} = \frac{x}{L_{0}}; \; \bar{t} = \frac{t}{(L/u_{0})}$

$\bullet$ Convective time scale

\begin{displaymath}\bar{\omega} = \frac{\omega}{u_{0}/L_{0}}\end{displaymath}

then $\frac{\partial \omega}{\partial \bar{t}} +\nabla \cdot (\vec{u} \omega) = Re^{-1} \nabla^{2} \omega$ in 2D, which is appropriate for modest to large Re, where

\begin{displaymath}Re = \frac{u_{0}L_{0}}{\nu} = \frac{{\rm inertial}}{{\rm viscous}}.\end{displaymath}

For $Re \rightarrow 0$, we introduce the diffusive time scale $t^{'} = \frac{t}{L^{2}/\nu}$, therefore the 2D-vorticity equation is: $\frac{\partial \omega }{\partial t} + Re \nabla \cdot (\vec{u}\omega) = \nabla^{2} \omega$, which is more appropriate for low Re (creeping) flows.

*Velocity-pressure: $\frac{\partial u}{\partial t} + u\cdot \nabla u = - \frac{\nabla p}{\rho} + \nu \nabla^{2} u$ (dimensional)

* $\frac{L_{0}}{u^{2}_{0}} \;\; \left (\frac{\partial\bar{u}}{\partial \bar{t}} + ...
...{0}L_{0}} \nabla ^{2} \bar{u}
= - \nabla \bar{p} + Re^{-1} \nabla^{2} \bar{u}$, where $\bar{p} = \frac{p}{\rho u^{2}_{0}}$

In summary: convective time units (fast): $\left (\frac{t}{L/u_{0}}\right )$, and difffusive time scale (slow): $\left ( \frac{t}{L^{2}/\nu}\right )$

IV. Boundary Conditions:
\begin{figure}
\centerline{\psfig{file=lec1.5.eps,width=3in}}
\end{figure}

$\bullet$ No-slip (Dirichlet) Us = Uw implies perfect momentum exhange of molecules with wall, molecules stick on the wall. Valid for continuum media. ( $\lambda \ll L_{0}$; the mean free path (in gases) much less than characteristic length scale)

Define $K_{n} \equiv \frac{\lambda}{L}$, Knudsen number

If Kn < 10-3, continuum $\Rightarrow$Direct Methods, (e.g., finite differences, finite elements, spectal methods). However, use Monte Carlo, Boltzmann for Kn > 10-3; rarefied allows slip and thus, model b.c. as

\begin{displaymath}U_{s} -U_{w} = K_{n} \frac{\partial U_{s}}{\partial \hat{n}}.\end{displaymath}



Applications: high-altitude, high-speed flows $(\lambda$ Large), micro-mechanis (L small) $\sim 1\mu m.$ ($\lambda =$ 65 nm at STP). $\bullet$ No-penetration (Dirichlet) For viscous or inviscid flows, in the norml direction we always have

\begin{displaymath}U_s \cdot \hat{u} = U_w \cdot \hat{n}\end{displaymath}

$\bullet$ Flux (Neumann b.c.)

\begin{displaymath}\underbrace{ \frac{\partial \theta}{\partial n} }_{\nabla \theta\cdot \hat{n}} = \underbrace{g(x,y,z,t)}_{{\rm known}}\end{displaymath}


\begin{figure}
\centerline{\psfig{file=lec1.6.eps,height=1in}}
\end{figure}

$\bullet$ Robin b.c. - (heat transfer; mass transfer) $k \frac{\partial\theta}{\partial n} = h (\theta-\theta_{\infty})$, where h = heat transfer coefficient.

$\bullet$ Outflow boundary conditions: Flow past an airfoil:
\begin{figure}
\centerline{\psfig{file=lec1.7.eps,height=2in}}
\end{figure}

outflow:
\begin{figure}
\centerline{\psfig{file=lec1.8.eps,height=1in}}
\end{figure}

NOTE: More difficult for incompressible than compressible where use of characteristics may work (not rigorous...)


Special Cases

2D Flow (does it exist?)
\begin{figure}
\centerline{\psfig{file=lec1.9.eps,height=2in}}
\end{figure}

$\bullet$ 2D geometry does not guarantee 2D flow. 3D for Re > Rec, bifurcation. Instabilities (bifurcations) may develop. Vorticity $\vec{\omega}_{z} = \hat{k} \nabla \times \vec{u}$, perpendicular to 2D plane in 2D flow. Stretching term $(\vec{\omega} \cdot \nabla) \vec{\omega} \equiv 0$ in 2D thus,

\begin{displaymath}\frac{\partial \omega_{z}}{\partial t} + (\vec{u} \cdot \nabla ) \omega_{z} = \nu \nabla^{2} \omega_{z}\end{displaymath}

Define $\psi$: (stream function) 4th order for $\psi$ (3rd for BL): $\vec{u} = -\nabla \times (\psi \hat{k})$, and thus $u=- \frac{\partial \psi}{\partial y} ; v = \frac{\partial \psi}{\partial x}$ Continuity: $\nabla \cdot \vec{u} \equiv 0$ is satisfied automatically and $\nabla^{2} \psi = - \omega$, expresses a kinematic constraint (it is not a time evolution equation).

For $\nu = 0\colon \; \frac{\partial \omega}{\partial t} + \left ( - \frac{\partial ...
...ght ) = 0
\Rightarrow \frac{\partial \omega}{\partial t} + J(\psi, \omega) = 0$

$\bullet$ steady state $\Rightarrow J(\psi, \omega) = 0 \Rightarrow \omega = f_{n} (\psi)$ that is, in inviscid 2D steady flow, vorticity is constant along steamlines. Streamlines $\equiv$ isovorticity lines

* Boundary Conditions: Too many for $\psi$/Thom's b.c. None for $\omega$ - create b.c. through local expansions and use the excessive $\psi$ b.c.

$\bullet$ Advantage: Incompressibility, the main headache in incompressible formulations, is honored automatically within discretization error. (Very popular method in the seventies.)

Stokes (creeping) flow:

\begin{displaymath}\left ( \vec{u}\cdot \nabla \right ) \vec{u} \ll \nu \nabla^{2} \vec{u},\end{displaymath}

i.e. very viscous flows $\rightarrow$ linear equations.

Applications: dispersion, combustion, particle flows small size $\rightarrow Re \ll 1\rightarrow$ approximate as Stokes flow.

parabolic/elliptic problems.
*
Stokes drag of a sphere : $D=6 \pi R \mu U_{\infty}$
\begin{figure}
\centerline{\psfig{file=lec1.10.eps,height=1in}}
\end{figure}

This is the oldest known analytical solution for a creeping flow (Stokes, 1851)
*
Here, we obtain symmetric solutions if geometry is symmetric. Contrast with symmetry-breaking bifurcations for finite Re.
Other complex shapes (particles): Boundary element methods

Inviscid flow:

\begin{displaymath}\nu \nabla^{2} \vec{u} \ll (\vec{u}\cdot \nabla)\vec{u}, \; {\rm for}\; Re \gg 1\end{displaymath}

e.g. high-speed flows
$\bullet$
hyperbolic equations
$\bullet$
There is some debate in the current literataure as to whether there exist singular solutions in incompressible flows, i.e., $\omega \sim (t-t_{0})^{m}, \; m < 0$ for t0 = finite.
$\bullet$
Standard vortex methods (Chorin/Leonard) work well for simulations of inviscid flows.
V. Potential Flow: $\vec{\omega} = 0$, $\vec{u} = \nabla \phi$ /incompressible, inviscid, irrotational flow, purely elliptic

\begin{eqnarray*}\nabla^{2} \psi &= &0\\
\nabla^{2} \phi &= & 0\; \mbox{{\rm with B.C.}}
\end{eqnarray*}


does it exist?
\begin{figure}
\centerline{\psfig{file=lec1.11.eps,height=1in}}
\end{figure}

at sufficiently large distances $u\rightarrow u_{\infty}$ in freestream and $\omega \simeq 0$ with $Re \gg 1$

$\bullet$ Compressible:

\begin{displaymath}\left ( 1 - \frac{u^{2}}{c_{0}} \right ) \frac{\partial \phi}...
...^{2}}{c_0}\right ) \frac{\partial^{2} \phi}{\partial y^{2}} = 0\end{displaymath}

+ energy equation

\begin{displaymath}(\vec{u} = \nabla \phi)\end{displaymath}

unknowns, $(\phi, c_0).$ Boundary Layer Flow
\begin{figure}
\centerline{\psfig{file=lec1.12.eps,height=1in}}
\end{figure}

neglect streamwise diffusion compared to normal (cross) diffusion $\delta /x \ll 1$, e.g.

\begin{displaymath}\frac{\partial u}{\partial t} + u \frac{\partial u}{\partial ...
...tial p}{\partial x} + \nu \frac{\partial ^{2}u}{\partial y^{2}}\end{displaymath}

great simplification, computation complexity in y-only; smaller domain.

*sometimes called the thin shear layer approximation
\begin{figure}
\centerline{\psfig{file=lec1.13.eps,height=2in}}
\end{figure}

In steady state we also call them: Parapolized N-S if the predominant main flow direction equation is parabolic in $x \leftarrow$ plays the role of time $\rightarrow$ advance explicitly. VI. Mathematical Nature of the Flow Equations

The balance of various terms in Navier-Stokes describes dominant diffusion or convection fluxes and thus the mathematical character of N-S will be varying from parabolic to hyperbolic depending on the flow problem.

$\bullet$ unsteady N-S: parabolic-hyperbolic; steady: elliptic-hyperbolic. The divergence-free constraint imposes the hyperbolicity.

*Examples: Stokes equation/creeping flow

\begin{displaymath}\frac{-u^{2}_{0}T_{0}}{\nu} \frac{\partial u}{\partial t} + \nabla^{2} u=Re \left ( \frac{\partial p}{\partial x}\right )\end{displaymath}

$\bullet$ steady state $\longrightarrow$ purely elliptic (for fixed $\frac{\partial p}{\partial x}$). $\bullet$ unsteady $\longrightarrow$ parabolic.

Outside the B.L. for $\nu = 0$ (high-speed flows) the governing equations are:

\begin{displaymath}\left \{ \begin{array}{lll}
\displaystyle{\frac{\partial u}{...
...frac{1}{\rho} \frac{\partial p}{\partial y}}\end{array}\right .\end{displaymath}

essentially a hyperbolic equation in space and time describing a propagation phenomenon. Inside the B.L. the governing equations are parabolic/hyperbolic.
PDEs of 2 $^{{\rm nd}}$-order (Review)

\begin{displaymath}a \frac{\partial ^{2}\phi}{\partial x ^{2}} + 2b \frac{\parti...
...ial ^{2} \phi}{\partial y^{2}} = 0; \; {\rm {\bf quasi-linear}}\end{displaymath}

Where $a,b,c = {\cal F} \left (x,y,\phi, \frac{\partial \phi}{\partial x}, \frac{\partial \phi}{\partial y}\right )$ (not 2nd derivatives). For example, $u = \frac{\partial \phi}{\partial x}; v = \frac{\partial \phi}{\partial y}$, where $\phi$ is the potential.

\begin{displaymath}\left \{ \begin{array}{ll}
\displaystyle{a \frac{\partial u}...
...rtial y} \left ( \begin{array}{cc}u\\ v\end{array} \right ) = 0\end{displaymath}




\begin{displaymath}\Rightarrow A_{1} \frac{\partial U}{\partial x} + A_{2} \frac{\partial U}{\partial y} = 0\end{displaymath}

$\bullet$ Solution: wave of the form $U = \hat{U} e^{i\vec{\eta}\cdot \vec{x}}, \vec{\eta} = (\eta_{x}, \eta_{y})$, $\vec{\eta} = $ propagation vector. Therefore, by substitution

\begin{displaymath}(A_{1} \eta_{x} + A_{2} \eta_{y} )\hat{U} = 0 \Rightarrow \det \vert A_{1} \eta_{x} + A_{2} \eta_{y}\vert=0\end{displaymath}


\begin{displaymath}\left \vert \begin{array}{ll}
a\eta_{x} + 2b\eta_{y} & c\eta...
...ht )^{2} + 2b \left ( \frac{\eta_{x}}{\eta_{y}}\right ) + c = 0\end{displaymath}


\begin{displaymath}\begin{array}{ll}
\mbox{{\rm Wave-like solution}}\\
\mbox...
...\\
\mbox{{\rm parabolic}} & b^{2}-ac = 0
\end{array}\right .\end{displaymath}

Example: Transonic-Inviscid Flow/Steady
\begin{figure}
\centerline{\psfig{file=lec1.14.eps,width=3in}}
\end{figure}

Local Mach $\left (\frac{u}{c_{0}}\right )$ variation; c0 = speed of sound

\begin{displaymath}\left (\underbrace{1-\frac{u^{2}}{c_{0}^{2}}}_{a}\right ) \fr...
...}^{2}}}_{c}\right ) \frac{\partial^{2}\phi}{\partial y^{2}} = 0\end{displaymath}

Determinant:

\begin{displaymath}b^{2}-ac = \frac{u^{2}+v^{2}}{c^{2}_{0}}-1 = M^{2}-1\end{displaymath}


\begin{displaymath}\begin{array}{ll}
\bullet M < 1, \;\; \mbox{{\rm subsonic fl...
...rm supersonic flow}} \rightarrow \; {\rm hyperbolic}\end{array}\end{displaymath}

This mixed nature of the transonic potential equation has been a great challenge in CFD since the transition line between the subsonic and the supersonic regions is part of the solution. Note that for $M = 1 \rightarrow$ parabolic.

$\bullet$ Non-linear wave propagation (general model) $U=\hat{U} e^{iS(x,y)}; S (x,y)$ is the wave characteristic surface, if real $\Rightarrow$ hyperbolic system. We thus consider, $\det (A_{1}S_{x} + A_{2} S_{y})=0$. If we interpret $\vec{\eta} = \nabla S$, normal to surface = direction of propagation, then $\Rightarrow U = \hat{U} e^{i(\vec{x}\cdot \nabla S)} = \hat{U} e^{i(xS_{x}+yS_{y})}$

Then the tangent vectors of the surface are defined as:

\begin{displaymath}dS = \vec{\nabla} S\cdot d\vec{x} = \frac{\partial S}{\partial x} dx + \frac{\partial S}{\partial y} dy = 0\end{displaymath}

Thus, the direction of the chacteristic surface is given by: $\frac{dy}{dx}= - \frac{S_{x}}{S_{y}} = - \frac{\eta_{x}}{\eta_{y}}$
\begin{figure}
\centerline{\psfig{file=lec1.15.eps,height=1in}}
\end{figure}

Note: The surface S(x,y) = const is a wavefront surface that separates the points already influenced by the propagating disturbance from the points not yet reached by the wave. Example: Potential flow for a thin airfoil:

\begin{displaymath}(1-M^{2}_{\infty} ) \frac{\partial^{2}\phi}{\partial x^{2}} + \frac{\partial ^{2} \phi}{\partial y^{2}} = 0\end{displaymath}


\begin{displaymath}\det = 0\Rightarrow a\left ( \frac{\eta_{x}}{\eta_{y}}\right ...
...arrow \frac{\eta_{y}}{\eta_{x}} = \pm \sqrt{M^{2}_{\infty} - 1}\end{displaymath}

Directon of Characteristic
\begin{figure}
\centerline{\psfig{file=lec1ab.eps,width=3in}}
\end{figure}


\begin{displaymath}\frac{\eta_{x}}{\eta_{y}} = \mp \frac{1}{\sqrt{M^{2}_{\infty} -1}} = \tan \mu\end{displaymath}


\begin{displaymath}\Rightarrow \sin \mu = \frac{1}{M_{\infty}}, \;\; \mbox{{\rm Mach lines}}\end{displaymath}



*System of $1^{{\rm st}}$-order equations/Generalization e.g. inviscid flows conservation (flux) form:

\begin{displaymath}\frac{\partial}{\partial x^{k}} F^{k}_{i} = Q_{i}, \begin{array}{ll}i=1,\ldots n\\
k=1, \ldots m\end{array}\end{displaymath}

$\rightarrow$ Quasi-linear form via Jacobian:

\begin{displaymath}A^{k}_{ij} = \frac{\partial F^{k}_{i}}{\partial u_{j}}\end{displaymath}


\begin{displaymath}\Rightarrow A^{k}_{ij} \frac{\partial u_{j}}{\partial x^{k}} ...
...}
i,j = 1, \ldots, n\\
k = 1, \ldots , m
\end{array}\right.\end{displaymath}

matrix form: (summation)

\begin{displaymath}A^{k} \frac{\partial U}{\partial x^{k}}= 0, \; k=1,\ldots, m\end{displaymath}

Ak, Q = F(xk,U) do not depend on derivatives of U!! A plane wave solution will exist with propagating direction $\vec{\eta}$ if $[A^{k}\eta_{k}]\hat{U} = 0 \Rightarrow \det (A^{k}\eta_{k})=0$, where $\eta_{k} =$ components of $\vec{\eta}$. The solutions $\vec{\eta} = [\eta_{k}], k=1,=\ldots, m$ are (n) altogether (equal to # of 1st order eqn) $\Rightarrow$ define characteristic surfaces from $\eta_{k} = \partial S/\partial x^{k} \Rightarrow $ at most (n) characteristic surfaces.

Definition:
Example: 2 equations:

\begin{displaymath}\left \{ \begin{array}{lll}
\displaystyle{a\frac{\partial u}...
...} + d \frac{\partial u}{\partial y} = F_{2}}\end{array}\right .\end{displaymath}


\begin{displaymath}\Rightarrow \left \vert \underbrace{\begin{array}{ll} a & 0\\...
... ) = \left ( \begin{array}{ll} F_{1}\\ F_{2}\end{array}\right )\end{displaymath}


\begin{displaymath}\vert A^{1}\underbrace{\eta_{1}}_{\eta_{x}} + A^{2} \eta_{2} ...
...\\
d & \frac{b\eta_{x}}{\eta_{y}}\end{array} \right \vert = 0\end{displaymath}


\begin{displaymath}\Rightarrow \left ( \frac{\eta_{x}}{\eta_{y}}\right )^{2} = \frac{cd}{ab}\end{displaymath}

Exercise: Show that the Cauchy-Riemann equations are of elliptic type

\begin{displaymath}C-R\left ( \begin{array}{ll}
\displaystyle{\frac{\partial u}...
...tial x} - \frac{\partial u}{\partial y} = 0}\end{array}\right )\end{displaymath}

Example: 2D Stationary shallow-water equations

\begin{displaymath}\left \{ \begin{array}{llll}\displaystyle{u \frac{\partial h}...
...{ccc}h\\ u\\ v\end{array}\right \} \; \mbox{{\rm state vector}}\end{displaymath}


\begin{displaymath}\left \vert\underbrace{\begin{array}{ccc}u & h & 0\\ g & u & ...
...= 0;\;\; {\rm where}\; \lambda \equiv \frac{\eta_{x}}{\eta_{y}}\end{displaymath}




\begin{displaymath}\left \vert \begin{array}{ccc}
u \lambda + v & h\lambda & h\...
...da + v & 0\\
g & 0 & u\lambda + v\end{array} \right \vert = 0\end{displaymath}

where


\begin{displaymath}\lambda^{(1)} = - \frac{v}{u}; \;\;\lambda^{(2),(3)} = \frac{-uv \pm \sqrt{u^{2}+v^{2} - gh}}{u^{2} - gh}\end{displaymath}

$\sqrt{gh} \approx$ ``sonic'' speed, $u^{2} + v^{2} > gh \Rightarrow$ hyperbolic

\begin{displaymath}\Rightarrow \; \mbox{{\rm else hybrid}}\end{displaymath}

$\bullet$ Characteristic surface $S^{1} \colon \vec{\eta}^{(y)} \cdot \vec{v} \Rightarrow \eta_{x} \cdot u + \eta_{y} v = (-v)u+(u)v=0$ $\Rightarrow$ characteristic surface $\equiv$ streamline surface.



Example: Inviscid, Incompressible Equations (Euler)/steady Define $U= \left ( \begin{array}{cc} u\\ v\\ p\end{array}\right )$ then

\begin{displaymath}\left [ \begin{array}{ccc}
u & 0 & 1/\rho\\ 0 & u & 0\\ 1 &0...
... 0 & 1 & 0\end{array}\right ] \frac{\partial U}{\partial y} = 0\end{displaymath}

Thus,

\begin{displaymath}\det \left [ \begin{array}{ccc} u + v \lambda & 0 & 1/\rho\\ ...
...ambda & \lambda/\rho\\
1 & \lambda & 0\end{array}\right ] = 0\end{displaymath}


\begin{displaymath}\Rightarrow (u+v\lambda)(\lambda^{2} + 1) = 0 \Rightarrow \be...
...m real}\\
\lambda_{2,3} = \pm i,\; {\rm imaginary}\end{array}\end{displaymath}

Threfore, this system is hybrid since it has mixed eigenvalues.

*Time-dependent flows We single out time from the system, e.g. xm = t,

\begin{displaymath}\Rightarrow \frac{\partial U}{\partial t} + A^{k} \frac{\part...
...; \underbrace{k=1,\ldots, m-1}_{\mbox{{\rm spatial variables}}}\end{displaymath}


\begin{displaymath}\det [\eta_{t} + \underbrace{A^{k}\eta_{k}}_{K_{ij}}]=0\end{displaymath}

equivalent to eigenvalue problem

\begin{displaymath}\det [K - \lambda I] = 0\end{displaymath}

where

\begin{displaymath}\eta_{t} = - \lambda\end{displaymath}

(compare with $U = \hat{U} e^{i(\vec{k}x-\omega t)} \rightarrow $ where, $\omega =$ frequency either real or complex, and $\vec{\eta}_{k} \rightarrow \vec{k}$

Example: 1D Time-dependent shallow water equations

\begin{displaymath}\left \{ \begin{array}{lll}
\displaystyle{\frac{\partial h}{...
...ght . ; U = \left \{ \begin{array}{c} h\\ u\end{array}\right \}\end{displaymath}


\begin{displaymath}\Rightarrow \frac{\partial U}{\partial t} + \left \vert \begi...
... u\\ \end{array} \right \vert \frac{\partial U}{\partial x} = 0\end{displaymath}


\begin{displaymath}\Rightarrow \left \vert \begin{array}{ccc}
u -\lambda & h\\ ...
... 0 \Rightarrow \lambda _{1,2,} = u \pm \sqrt{gh}\; \epsilon \Re\end{displaymath}


\begin{displaymath}\Rightarrow \; \mbox{{\rm hyperbolic system}}\end{displaymath}

Question: What is the 2D time-dependent shallow water? Domain of Dependence/Influence
\begin{figure}
\centerline{\psfig{file=lec4.9.eps,width=4in}}
\end{figure}

$\Theta_{x} + \Theta_{y} = 0$ $\Theta_{x} = \Theta_{yy}$ $ \Theta_{xx} + \Theta_{yy} = q$
$\bullet$ hyperbolic $\bullet$ parabolic $\bullet$ elliptic
$\bullet$ IVP $\bullet$ IVP/BVD $\bullet$ BVP
Computational Complexity of Navier-Stokes
*At high Re Kolmogorov's scaling theory; also Foias's argument
\begin{figure}
\centerline{\psfig{file=lec1.17.eps,width=3in}}
\end{figure}

Assume that L = size of a large eddy and that (urms) represents $\{rms\}$ value. Also $\epsilon$ is the dissipation $(\epsilon \sim u^{3}/L)$ and $\eta =$ Kolmogorov's scale, which is the smallest scale of turbulence.

1D: disparity of scales Number of Degrees of freedom $\displaystyle{\sim \frac{L}{\eta} Re^{3/4}}$ $Re_{\eta} = \displaystyle{\frac{\eta u_{rms}}{\nu}} = 1$
3D: scales carry energy $\sim Re^{9/4}$ $\bullet$ Time scales: fast: $\displaystyle{\frac{\eta}{v_{rms}}}$; slow: $\displaystyle{\frac{L}{v_{rms}}}$

$\bullet$ Explicit integration in time dictates that we honor the CFL condition, i.e.,

\begin{eqnarray*}\Delta t & < & CFL< {\cal O} (1)\\
& \Rightarrow & \Delta t < \frac{\eta}{u_{rms}}
\end{eqnarray*}


$\bullet$ Need to integrate for long time $T = \frac{L}{u_{rms}}$ (one large eddy-turnover time). Thus, number of time steps:

\begin{displaymath}\frac{T}{\Delta t} \sim \frac{L}{\eta} = Re^{3/4}\end{displaymath}

Thus total work $\sim (Re^{9/4}) (Re^{3/4}) = Re^{3}$, so increasing Re by a factor of 2 leads to an order of magnitude increase in the CPU requirement.

Recommended Reading:
(1)
``Computer Experiments in Fluid Dynamics,'' Harlow, F.H. and Fromm, J.E., Scientific American, vol. 212, No. 3, pp 104-110.
(2)
``Nodes, Modes and Flow Codes'', Karniadakis, G.E. and Orszag, S.A., Physics Today, March 1993, pp 34-42.
(3)
G. Batchelor's book: Derivation of the Navier-Stokes Equations and Conservation Laws
(4)
P-L. Lions, Mathematical Topics in Fluid Mechanics, Oxford Science Publications, 1996.
(5)
P. Constantin and C. Foias, Navier-Stokes Equations, University of Chicago Press, 1988.
(6)
V.A. Solonnikov and A.V. Kazhmikov, ``Existence theorems for the equations of a compressible viscous fluid,'' Ann. Rev. Fluid Mech., vol. 13, p. 79, 1982.